首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Aqueous solutions of indium(III) undergo 1e reductions by Sm(II) (E0-1.55 V) and by Yb(II) (−1.05 V), but not by U(III) (−0.52 V). The facile and irreversible reduction, by In+ aq, of Ru(NH3)6 3+ (E0+0.067 V) at reagent concentrations near 10−3 M implies a potential more negative than −0.23 V for In(II,I) and, in conjunction with the known value of −0.44 V for In(III,I), a complementary potential less negative than −0.65 V for In(III,II). These observations, taken together, support le indium potentials EII,I o−0.2 V and EIII,II 0−0.6 V.  相似文献   

2.
It has been found that the modified Zhuravlev equation, [(1−α)−1/3−1]2=ktn, which describes the kinetics of oxidation of V2O4 and V6O13 in the temperature range 820–900 K and in the oxygen pressure range 1.0–20 kPa, can be derived via the assumption that the changes in the observed activation energy result from the changing contributions of the two diffusion processes controlling the reaction rate. The values of the observed activation energy are in the range 160–175 kJ mol−1 for V2O4 and 188–201 kJ mol−1 for V6O13 in the scope of the experimental oxygen pressures and temperatures and conversion degrees of 0.1–0.9. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

3.
The influence of TiOSO4 and free sulphuric acid concentrations in the starting solution on the degree of titanyl sulphate conversion to hydrated titanium dioxide and post-hydrolytic sulphuric acid was studied. Titanyl sulphate solution, an intermediate product in the commercial preparation of titanium dioxide pigments by sulphate route, was used. It was found that the degree of hydrolysis markedly depends on the studied parameters. The lower was the content of TiOSO4 in the starting solution, the higher conversion was achieved. The degree of hydrolysis at the final stage varied between 81 % (420 g TiOSO4 dm−3, 216 g H2SO4 dm−3) and 92 % (300 g TiOSO4 dm−3, 216 g H2SO4 dm−3). The same relation was obtained when changing the concentration of free H2SO4 in the starting solution. The degree of hydrolysis at the final stage varied between 49 % (261 g H2SO4 dm−3, 340 g TiOSO4 dm−3) and 96 % (136 g H2SO4 dm−3, 340 g TiOSO4 dm−3). The particle size of the obtained hydrated titanium dioxide (HTD) also depends on the initial solution composition. Presented at the 34th International Conference of the Slovak Society of Chemical Engineering, Tatranské Matliare, 21–25 May 2007.  相似文献   

4.
The bis(chelated) complex of CrV(0) derived from the dianion (L2 ) of 2-ethyl-2-hydroxybutanoic acid is readily reduced to a bis(chelate of CrIII, featuring the monoanion (LH) [Cr V(0)(L2−)2]+4H++H2O+2e→[CrIII(OH2)2(LH 2]+ of this acid. Potentials estimated by Ghosh in 1993 for this 2e change, E0 (pH 0) 1.32 V, Eeff (pH 3.3) 0.93 V, are in accord with the nearly irreversible reductions of the Cr(V) species (in 1∶1 ligand buffer) by Fe2+, V02+, IrCl6 3 and I, whereas lower values reported by Bose in 1996, E0 (pH 0) 0.84 V, Eeff (pH 3.3) 0.45 V, are potentiometrically inconsistent with these conversions. A similar discrepancy is noted for potentials for Cr(V,IV) estimated in 1996, E0 (pH 0) 0.84 V, Eeff (pH 3.3) 0.46 V, which, wholly contrary to observation, predict that the reductions of excess Cr(V) to CR(IV) by Fe2+, V02+, and I are thermodynamically disfavored.  相似文献   

5.
The reaction of 1-ethoxycarbonylmethyl-5,5,7,7-tetramethyl-2-oxo-tetrahydroimidazo[1,5-b]oxadiazol-6-oxyl with the weakly basic nucleophiles NaN3, NaCN, KF, KBr, KCl and NaNO2 has been studied. It was shown for the first time that, as in the case of NaOH and MeONa, the reaction occurs with opening of the oxadiazolone ring to form exo-N-substituted amidines. It was shown that the weakly basic nucleophiles readily react with substrates which contain a substituent sensitive to attack by such nucleophiles as NaOH or MeONa. The effect of the nature of the nucleophiles on the reaction course for opening of the oxadiazolone ring was also studied. It was found that the reactivity of the nucleophiles in DMSO changes in the series F > CN > N3 >NO2 > Cl > Br and qualitatively correlates with their basicities in this solvent. Examination of the effect of the ratio of the reagents on the degree of conversion of the starting oxadiazolone has shown that a quantity of nucleophiles less than one equivalent also allowed the cleavage reaction of the oxadiazolone heterocycle to go to completion through just increasing the reaction time. The experimental data obtained lends support to the proposed reaction scheme. Dedicated to Academician B. A. Trofimov in his 70th jubilee. Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 1, pp. 71–78, January, 2009.  相似文献   

6.
Summary An ion-chromatographic procedure is described for the determination of selenium (VI) at μg L−1 level in the presence of anions and heavy metal ions. Maximum permissible concentrations and effects from each interfering substance were investigated for the Se concentration range 12.5–1,000 μg L−1. The method, optimized for the detection of SeO 4 2− , gives results suitable for speciation analysis. Total selenium can be determined after complete conversion to selenate ion by oxidation with KMnO4. The detection limit of selenium is 4.8 μg L−1 (0.96 ng for 200 μL sample). Paper presented at the 41st Pittsburgh Conference, New York, March 5–9, 1990.  相似文献   

7.
The action of ozone on a suspension of Fe(III) hydroxide in alkaline solutions was studied by the spectrophotometric method. A partial conversion of Fe(III) to Fe(VI) is observed at Fe(III) concentrations exceeding 2 mmol l−1. The tenfold increase of the initial Fe(III) concentration raises the Fe(VI) yield by a factor of 2–3. The mechanism of the process includes the decomposition of ozone with the formation of ozonide ions, which oxidize Fe(III) up to Fe(IV), Fe(V), and Fe(VI) in parallel with their conversion to O2 and HO2. Fe(VIII) is not formed.  相似文献   

8.
The formation of neutral Tl2max = 390 nm) and Tl4max = 360 nm) clusters in dilute aqueous solutions of Tl2SO4 containing formate ions was found by pulse radiolysis. The rate constants for the recombination of Tl0 atoms and Tl2 clusters are equal to 1.5·1010 L mol−1 s−1 and 1.0·1010Lmol−1 s−1 (±30%), respectively, and the extinction coefficient of Tl2 at 390 nm is −6.0·103 L mol−1 cm−1 Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2367–2369, December, 1999.  相似文献   

9.
The kinetics of the oxidative conversion of 1,3,7-trimethylxanthine upon treatment with hypochlorite ions (OCl) in aqueous medium at 283–298 K and pH 8.2 was studied. The reaction order with respect to each component was determined and proved to be 1. It was established that the temperature dependence of the reaction rate follows the Arrhenius equation. The activation parameters of the reaction were measured: E a = 33.58 kJ/mol, ΔH = 31.12 kJ/mol, ΔS = −170.02 J/(K mol), ΔG = 81.45 kJ/mol. The stoichiometry of the reaction was studied, and the chemistry of the oxidative conversion of caffeine treated with OCl is discussed.  相似文献   

10.
The adsorption and activation of NO molecules on Cu-ZSM-5 catalysts with different Cu/Al and Si/Al ratios (from 0.05 to 1.4 and from 17 to 45, respectively) subjected to different pretreatment was studied by ultraviolet-visible diffuse reflectance (UV-Vis DR). It was found that the amount of chemisorbed NO and the catalyst activity in NO decomposition increased with an increase in the Cu/Al ratio to 0.35–0.40. The intensity of absorption bands at 18400 and 25600 cm−1 in the UV-Vis DR spectra increased symbatically. It was hypothesized that the adsorption of NO occurs at Cu+ ions localized in chain copper oxide structures with the formation of mono- and dinitrosyl Cu(I) complexes, and this process is accompanied by the Cu2+...Cu+ intervalence transfer band in the region of 18400 cm−1. The low-temperature activation of NO occurs through the conversion of the dinitrosyl Cu(I) complex into the π-radical anion (N2O2) stabilized at the Cu2+ ion of the chain structure, [Cu2+-cis-(N2O2)], by electron transfer from the Cu+ ion to the cis dimer (NO)2. This complex corresponds to the L → M charge transfer band in the region of 25600 cm−1. The subsequent destruction of the complex [Cu2+-cis-(N2O2)] at temperatures of 150–300°C leads to the release of N2O and the formation of the complex [Cu2+O], which further participates in the formation of the nitrite-nitrate complexes [Cu2+(NO2)], [Cu2+(NO)(NO2)], and [Cu2+(NO3)] and NO decomposition products.  相似文献   

11.
Low-energy CAD product-ion spectra of various molecular species of phosphatidylserine (PS) in the forms of [M−H] and [M−2H+Alk] in the negative-ion mode, as well as in the forms of [M+H]+, [M+Alk]+, [M−H+2Alk]+, and [M−2H+3Alk]+ (where Alk=Li, Na) in the positive-ion mode contain rich fragment ions that are applicable for structural determination. Following CAD, the [M−H] ion of PS undergoes dissociation to eliminate the serine moiety (loss of C3H5NO2) to give a [M−H−87] ion, which equals to the [M−H] ion of a phoshatidic acid (PA) and give rise to a MS3-spectrum that is identical to the MS2-spectrum of PA. The major fragmentation process for the [M−2H+Alk] ion of PS arises from primary loss of 87 to give rise to a [M−2H+Alk−87] ion, followed by loss of fatty acid substituents as acids (RxCO2H, x=1,2) or as alkali salts (e. g., RxCO2Li, x=1,2). These fragmentations result in a greater abundance of [M−2H+Alk−87−R2CO2H] than [M−2H+Alk−87−R1CO2H] and a greater abundance of [M−2H+Alk−87−R2CO2Li] than [M−2H+Alk−87−R1CO2Li]; while further dissociation of the [M−2H+Alk−87−R2(or 1)CO2Li] ions gives a preferential formation of the carboxylate anion at sn-1 (R1CO2) over that at sn-2 (R2CO2). Other major fragmentation process arises from differential loss of the fatty acid substituents as ketenes (loss of Rx′CH=CO, x=1,2). This results in a more prominent [M−2H+Alk−R2′CH=CO] ion than [M−2H+Alk−R1′CH=CO] ion. Ions informative for structural characterization of PS are of low abundance in the MS2-spectra of both the [M+H]+ and the [M+Alk]+ ions, but are abundant in the MS3-spectra. The MS2-spectrum of the [M+Alk]+ ion contains a unique ion corresponding to internal loss of a phosphate group probably via the fragmentation processes involving rearrangement steps. The [M−H+2Alk]+ ion of PS yields a major [M−H+2Alk−87]+ ion, which is equivalent to an alkali adduct ion of a monoalkali salt of PA and gives rise to a greater abundance of [M−H+2Alk−87−R1CO2H]+ than [M−H+2Alk−87−R2CO2H]+. Similarly, the [M−2H+3Alk]+ ion of PS also yields a prominent [M−2H+3Alk−87]+ ion, which undergoes consecutive dissociation processes that involve differential losses of the two fatty acyl substituents. Because all of the above tandem mass spectra contain several sets of ion pairs involving differential losses of the fatty acid substituents as ketenes or as free fatty acids, the identities of the fatty acyl substituents and their positions on the glycerol backbone can be easily assigned by the drastic differences in the abundances of the ions in each pair.  相似文献   

12.
Solid electrolyte membrane reactors (SEMRs) have been used to both study and influence catalytic reaction rates. Methane coupling is the reaction most thoroughly and intensively studied in these membrane reactors. In the last 20 years, oxygen ion (O2−), proton (H+) and mixed (O2−-e, H+-e) conducting membranes have been tested in order to maximize the conversion of methane to C2 compounds. The present review contains the fundamental operating principles of the various SEMR types and their applications in this reaction. The difficulties that should be overcome in order to promote this SEMR process to an industrial scale are discussed.  相似文献   

13.
Kinetic studies on Li+ exchange between the cryptands C222 and C221, and γ-butyrolactone as solvent were performed as a function of ligand-to-metal ratio, temperature and pressure using 7Li NMR. The thermal rate and activation parameters are: C222: k 298 = (3.3 ± 0.8)×104 M−1 s−1, ΔH # = 35 ± 1 kJ mol−1 and ΔS # = −41 ± 3 J K−1 mol−1; C221: k 298 = 105 ± 32 M−1 s−1, ΔH # = 48 ± 1 kJ mol−1 and ΔS # = −45 ± 2 J K−1 mol−1. Temperature and pressure dependence measurements were performed in the presence of an excess of Li+. The influence of pressure on the exchange rate is insignificant for both ligands, such that the value of activation volume is around zero within the experimental error limits. The activation parameters obtained in this study indicate that the exchange of Li+ between solvated and chelated Li+ ions follows an associative interchange mechanism. Electronic Supplementary Material Supplementary material is available to authorised users in the online version of this article at . For Part I see: R. Puchta, M. Galle, N.J.R. van Hommes, E. Pasgreta and R. van Eldik: Inorg. Chem. 43, 8227 (2004).  相似文献   

14.
The rate constants for the reactions of positronium with I 3 , Br, and I in dimethylsulphoxide (DMSO) and for I 3 in water have been determined, and the orthopositronium lifetime variations have been used for studying polyhalide formation in the (I2+Br)/DMSO system. It is found that the formation of more than one polyhalide is needed to account for the data. On basis of this new evidence, previous results in methanol and in DMSO are re-examined. Besides the primary I2X (X: halide), I2X 2 2− complexes have to be invoked. Trial absorption measurements confirm the presence of solvolysis equilibria for I2 in DMSO. Taking these facts into account, much higher, albeit poorly defined, values than previously reported are found for the I2X formation constants. The reliable Kc values deduced are 28M−1, 0.57M−1 and 2.5M−1 respectively for I2Cl, I2Cl 2 2 and I2Br 2 2− in methanol, and 1.4M−1 and 0.8M−1, respectively for I2Cl 2 2− and I2Br 2 2− in DMSO.  相似文献   

15.
16.
Two multiamide calix[4]arenes (5, 6) were synthesized and characterized by NMR, MS and elemental analysis. The binding properties of receptors with some anions (π-O2NPhOPO32−, π-O2NPhO, H2PO4, Ac, Cl, Br and I) were studied by UV-Vis spectra. The results indicate that the tetraamide calix[4]arenes (5, 6) have a good selectivity to the anions containing aromatic ring (π-O2NPhOPO3 2−, π-O2NPhO). The 1 : 1 complexes between host and guest were formed through multiple hydrogen bonding and π-π interactions. The hosts 5 and 6 also show a definite binding ability for the anions (H2PO4, Ac, Cl) that have no ultraviolet absorption, which provides a simple method of spectrum detection for these anions.  相似文献   

17.
The kinetics of nucleophilic substitution of pyridine in bis-cationic [Pt(L)(py)]2+ complexes (L=SNS, NNN, NSN) [SNS=bis(methylthiomethyl)pyridine, NNN=bis(2-pyridylmethyl)amine, NSN=bis(2-pyridylmethyl)sulphide] by a series of nucleophiles (Cl, Br, I, N3, (C2H5)2S, NH3, thiourea (tu), NO2, C5H10NH, SeCN, SCN, CN when L=SNS; Cl, Br, I, N3, (C2H5)2S, SCN, NH3, NO2 when L=NNN; Br, N3, NO2, NH3, C5H10NH when L=NSN) have been measured in MeOH at 25 °C, μ =0.1 mol dm−3 (LiClO4 or LiCF3SO3). The logarithms of the second-order rate constants calculated at μ=0, log k° 2, do not follow the dependence upon the n° Pt scale. In particular, the reactivity of the biphilic reagents tu, SeCN, SCN and, to a lesser extent, NO 2, towards these doubly charged substrates is largely lower than expected on the basis of the n° Ptscale. There are good linear relationships between logk° 2 for the bis-cationic substrate [Pt(SNS)(py)]2+, chosen as the standard, and log k° 2 for the same reactions with [Pt(NNN)(py)]2+, [Pt(NSN)(py)]2+ and other double charged complexes previously studied. A new wide nucleophilicity scale based on [Pt(SNS)(py)]2+, that is appropriate to all the bis-cationic substrates, is here proposed  相似文献   

18.
Sodium diethyldithiocarbamate (DDTC-Na) was demonstrated to be a new colorimetric cyanide chemosensor by utilizing an indirect trick. First, some copper ions were added to the colorless aqueous solution of DDTC-Na. Then, the resultant brown solution was studied upon the addition of different anions, including Cl, I, IO3, SO42−, NO2, Br, H2PO4, F, SCN, HSO4, ClO4 and CN. It was observed by naked eyes that the brown solution changed to colorless immediately after the addition of the trace cyanide, but there were no changes towards other anions, making DDTC-Na a good selective cyanide chemosensor in pure water. Supported by the National Natural Science Foundation of China (Grant Nos. 20674059 & 20402011)  相似文献   

19.
In vivo imaging of reactive small molecule metabolites with high spatial resolution and specificity could give clues to understanding pathophysiology of various diseases. We herein applied time of flight-secondary ion mass spectrometry (TOF-SIMS) to newly developed silver-deposited plates that were stamped on mouse tissues, and succeeded in visualization of halide (Cl, Br, and I) and pseudohalide thiocyanate (SCN) anions, a class of substrates for neutrophils/eosinophil peroxidases to produce hypohalous acids (HOX/OX mixture; X: (pseudo)halides), as well as hydrogen sulfide (H2S). Forty-micrometer frozen mouse kidney sections on cover glasses were attached to 37 °C preheated silver-deposited plates and incubated at −10 °C for 1 h. After sputter cleaning to remove surface contaminants, the plates were analyzed by TOF-SIMS to identify distribution of Br, AgBr2, I, AgI2, SCN, as well as S2− and AgS as products of tissue-derived H2S. Br, AgBr2, I, and SCN anions were mainly distributed in core regions including the inner medulla and inner stripe of the outer medulla (except for I), rather than outer regions such as the cortex and outer stripe of the outer medulla. AgI2 anion was spread over the whole kidney, although its levels were relatively low. In contrast, S2− and AgS anions were mainly present in the outer regions. To our knowledge, this is the first imaging study to reveal the distribution of (pseudo)halides and H2S in animal tissue sections.  相似文献   

20.
Photocatalytic reduction of CO2 to HCOOH and HCHO was carried out in a Pt/CdS/RuO2 semiconductor particulate system using [RuIII(EDTA-H)H2O] complex as catalyst. Upon illumination at 505 nm (band gap energy of CdS), the system produced HCOOH and HCHO at rates equal to 3.05 × l0−2 Mh−1 and 2.0 × 10−2 M h−1, respectively. Trace amounts of CH2OH, CH4 and CO were also detected in the reaction vessel. Photobiological conversion of CO2 to formic acid was achieved by usingHalobacterium halobium MMT22 in aqueous solution at a rate equal to 0.45 M h−1. A one-and-half-fold increase in the rate of formation of formic acid was observed when the photobiological reduction of CO2 was performed in the presence of L-ascorbic acid as electron-donating agent and [RuIII(bipy)3]2+ as photosensitizer.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号